This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

A unified computational framework for modeling genome-wide nucleosome landscape

, and

Published 12 September 2018 © 2018 IOP Publishing Ltd
, , Citation Hu Jin et al 2018 Phys. Biol. 15 066011 DOI 10.1088/1478-3975/aadad2

1478-3975/15/6/066011

Abstract

Nucleosomes form the fundamental building blocks of eukaryotic chromatin, and previous attempts to understand the principles governing their genome-wide distribution have spurred much interest and debate in biology. In particular, the precise role of DNA sequence in shaping local chromatin structure has been controversial. This paper rigorously quantifies the contribution of hitherto-debated sequence features—including G+C content, 10.5 bp periodicity, and poly(dA:dT) tracts—to three distinct aspects of genome-wide nucleosome landscape: occupancy, translational positioning and rotational positioning. Our computational framework simultaneously learns nucleosome number and nucleosome-positioning energy from genome-wide nucleosome maps. In contrast to other previous studies, our model can predict both in vitro and in vivo nucleosome maps in Saccharomyces cerevisiae. We find that although G+C content is the primary determinant of MNase-derived nucleosome occupancy, MNase digestion biases may substantially influence this GC dependence. By contrast, poly(dA:dT) tracts are seen to deter nucleosome formation, regardless of the experimental method used. We further show that the 10.5 bp nucleotide periodicity facilitates rotational but not translational positioning. Applying our method to in vivo nucleosome maps demonstrates that, for a subset of genes, the regularly-spaced nucleosome arrays observed around transcription start sites can be partially recapitulated by DNA sequence alone. Finally, in vivo nucleosome occupancy derived from MNase-seq experiments around transcription termination sites can be mostly explained by the genomic sequence. Implications of these results and potential extensions of the proposed computational framework are discussed.

Export citation and abstract BibTeX RIS

1. Introduction

Eukaryotic DNA is tightly packaged inside the nucleus through a hierarchical structure of chromatin. The first level of compaction involves wrapping 147 base pairs (bps) of DNA around a histone octamer, forming the nucleosome that repeatedly occurs across the genome. The precise locations of nucleosomes on genomic DNA influence higher-order chromatin structure [1] and regulates diverse cellular processes by controlling local DNA accessibility [2]. Specifically, it is believed that most DNA-binding proteins preferentially bind the unprotected linker DNA between nucleosomes. Discovering the rules that govern genome-wide nucleosome landscape is thus a major step towards understanding nucleosome-mediated chromatin structure and regulatory processes taking place on the chromatin template.

Decades of research have demonstrated that no single mechanism can explain every aspect of the nucleosome landscape, and several competing ideas have been proposed to date. First, both in vitro [3, 4] and in vivo [58] studies have suggested that DNA sequence may play a role, since experimental evidence suggests that certain sequences either favor or disfavor nucleosome formation. Second, nucleosome locations can be actively modulated by trans-factors, including chromatin remodelers, transcription factors, and RNA polymerase [2]. Finally, statistical positioning arising from a sharp barrier and the steric exclusion between neighboring nucleosomes may account for a large portion of nucleosome arrays around transcription start/termination sites (TSS/TTS) [9, 10]. The relative contributions of these factors in shaping local chromatin structure remain unclear and, in fact, highly controversial [24, 11, 12]. In particular, to what extent the global nucleosome landscape is a priori programmed into the genomic sequence itself remains to be determined.

To date, various sequence features, such as the G+C content (GC), 10.5 bp nucleotide periodicity, and poly(dA:dT) tracts (polyA), have been shown to influence nucleosome formation; however, the genome-wide contributions of these individual features have not been completely characterized. Another impediment to understanding the determinants of genome-wide nucleosome landscape stems from the confusion about different technical concepts involved [13, 14]. That is, different sequence features may play different roles in affecting nucleosome occupancy—referring to the probability of a given genomic location being occupied by a nucleosome—versus nucleosome positioning—referring to the precise genomic locations of individual nucleosomes, which can be described by the start or center (dyad) locations of the nucleosomes [13, 14]. The distinction between these two concepts is often blurred in the literature. More precisely, the concept of nucleosome positioning can be further distinguished into translational and rotational positions. The translational nucleosome position refers to the location of the 147 bp DNA contacting the histone octamer, while the rotational nucleosome position refers to the rotational orientation of DNA double helix relative to the histone surface [13, 14]. Previous studies have focused on either a single sequence feature (e.g. [15] studies the role of 10.5 bp periodicity in genome-wide nucleosome occupancy and nucleosome positioning in synthetic DNA sequences) or a single aspect of nucleosome landscape (e.g. [16] studies the role of GC and 10.5 bp periodicity in nucleosome occupancy). By contrast, this paper presents a unified framework for quantitatively characterizing the role of several distinct sequence features (GC, 10.5 bp periodicity, and polyA) in nucleosome occupancy, translational nucleosome positioning, and rotational nucleosome positioning, both in vitro and in vivo. To avoid confusion in the rest of the paper, the term 'nucleosome landscape' will be used to refer to the general biological problem encompassing nucleosome occupancy, translational nucleosome positioning, and rotational nucleosome positioning.

The 'beads-on-a-string' structure of nucleosome-occupied DNA is reminiscent of the 'Tonks gas' model [17] of non-overlapping one-dimensional (1D) rods. Indeed, this model has been used to quantify the role of statistical positioning and explain the regularly-spaced nucleosome arrays observed around TSS and TTS [9, 10, 18, 19]. In order to further account for sequence effects, a 1D sequence-dependent binding energy can be introduced into the model. This resulting model of 1D hard rods subjected to an external potential has been widely applied to study the genome-wide distribution of nucleosomes [15, 16, 2026]. In particular, Locke et al consider the inverse problem of inferring the nucleosome-positioning energy from genome-wide nucleosome maps generated through high-throughput sequencing, regressing this energy as a linear function of DNA sequence content [16]. They conclude that GC is the primary determinant of nucleosome occupancy, with the 10.5 bp periodicity making little contribution, and that DNA sequence alone cannot create the nucleosome arrays observed around TSS and TTS in vivo. The work of Locke et al has provided refreshing perspectives in the field, and their approach has been extended in several subsequent studies [23, 24, 26]. However, we here show that the calculation performed by Locke et al incorrectly estimates the total number of nucleosomes in the genome (we refer to this number as the nucleosome number, section 2.1). We demonstrate that this problem ultimately leads to incorrect inference of the nucleosome-positioning energy, thereby confounding the downstream analysis of sequence effects (section 2.1).

The main difficulty in the approach of Locke et al arises from their normalization scheme for converting the raw counts of sequencing reads to well-defined probability values. In this paper, we first develop a cross-entropy method (CEM) that gets around this normalization problem and simultaneously learns nucleosome number as well as nucleosome-positioning energy from genome-wide nucleosome maps (section 2.2). Our CEM is inspired by similar optimization-based approaches [15, 25, 26], but contains no artificial components (as in [15, 26]) and provides a flexible framework that can incorporate various regulatory factors, such as chromatin remodelers, which may affect nucleosome occupancy and positioning (section 4).

We validate our method using several independent nucleosome maps in S. cerevisiae and demonstrate that it consistently outperforms the method of Locke et al in predicting nucleosome occupancy, benefiting from the ability to learn the correct nucleosome number, even though CEM contains much fewer free parameters (section 3.1). We then present a comprehensive quantification of the influence of GC, 10.5 bp periodicity, and polyA on nucleosome occupancy and positioning. Importantly, we show that the GC dependence of MNase-derived nucleosome occupancy is substantially skewed by MNase digestion biases [27, 28], while polyA excludes nucleosome occupancy regardless of the experimental method used (sections 3.2 and 3.3). We also establish that the 10.5 bp periodicity, although not significantly affecting nucleosome occupancy, facilitates rotational but not translational positioning [29] (section 3.4). Further application of CEM to in vivo nucleosome maps shows that the oft-observed regularly spaced nucleosome array at TSS can be partially recapitulated by DNA sequence alone for a subset of genes, in sharp contrast to the conclusion of Locke et al (sections 3.5 and 3.6). We argue that this discrepancy stems from the previous under-estimation of nucleosome number. Finally, we show that MNase-derived nucleosome occupancy around TTS in S. cerevisiae is mostly encoded in the genomic sequence (section 3.7). Our work thus rigorously quantifies of the effect of DNA sequence on nucleosome formation and highlights the importance of correctly estimating nucleosome number in modeling in vivo and in vitro nucleosome landscapes.

2. Methods

2.1. Evaluation of the effect of estimated nucleosome number on model predictions

Nucleosome distribution can be modeled as the hard-core Tonks gas of 1D rods [16, 23, 24, 30] (supplementary material (stacks.iop.org/PhysBio/15/066011/mmedia)4, section 1.1). In the simplest form, nucleosomes are assumed to occupy a fixed number a of base pairs (bps) along DNA of length L and cannot overlap with each other due to steric exclusion. This steric exclusion is realized by imposing hard-core interaction between neighboring nucleosomes. The total energy of the system then involves only the sequence-dependent nucleosome-positioning energy u due to histone-DNA interactions, with each nucleosome occupying bps $i, i+1, \ldots, i+a-1$ contributing an energy term $u(i)$ at location i. Given a fixed form of u, the statistical mechanics of nucleosome landscape concerns solving for the one-particle distribution function n1(i), the probability of a nucleosome starting at bp i, and the nucleosome occupancy $O(i) = \sum_{j = i-a+1}^{i}n_1(\,j)$ , the probability of bp i being occupied by some nucleosome. This problem of calculating n1 and O from known u is termed the direct problem, which can be solved using techniques from statistical mechanics. However, the inverse problem of calculating u from known n1 and O is more difficult and also more pertinent in biology, since n1 and O can be directly linked to experimental data, as described below.

This 1D hard-rod model and its variants have been widely applied to study various questions associated with genome-wide nucleosome landscapes [9, 10, 15, 16, 2026]. In particular, Locke et al have used this approach to investigate how DNA sequence regulates nucleosome occupancy [16]. Specifically, n1 and O are first obtained from high-throughput sequencing data of genome-wide nucleosomes; then, the nucleosome-positioning energy u (or rather, $\frac{u - \mu}{k_B T}$ , where μ denotes chemical potential, kB the Boltzmann constant, and T temperature) is calculated by solving the inverse problem in the Grand Canonical Ensemble. To simplify notation, we use u to denote the dimensionless quantity $\frac{u - \mu}{k_B T}$ throughout the paper. Finally, linear models on DNA sequences are fitted to the calculated u to study how different sequence features influence the nucleosome energetics (supplementary material, section 1.2). We here describe an important problem associated with their very first step of estimating n1 and O from experimental data and demonstrate how this problem might lead to incorrect conclusions about the sequence dependence of nucleosome landscape.

Raw data from a high-throughput genome-wide nucleosome mapping experiment, such as MNase-seq, yield the nucleosome count $\tilde{n}_1(i)$ and the observed nucleosome occupancy $\tilde{O}(i) = \sum_{j = i-a+1}^i \tilde{n}_1(\,j)$ , where $\tilde{n}_1(\,j)$ is the number of sequenced nucleosomes starting at bp j and $\tilde{O}(i)$ is the the number of sequenced nucleosomes covering bp i. Under the assumption that the true one-particle distribution function is proportional to the observed nucleosome count, i.e. $n_1 \propto \tilde{n}_1$ , n1 may be estimated by properly scaling $\tilde{n}_1$ . In Locke et al, n1 is estimated as

Equation (1)

such that the obtained n1 and O are both bounded between 0 and 1 and, thus, have probabilistic interpretations.

We note several problems associated with this estimation. First, it artificially sets the maximum value of O to 1; in other words, at least one bp in the genome is assumed to have probability 1 of being occupied by some nucleosome, but this assumption is not necessarily true. Second, the normalization coefficient $\max_j{\tilde{O}(\,j)}$ is determined by read counts in only a very small region of the genome and, thus, may be sensitive to experimental noise and bias. Most importantly, equation (1) does not guarantee the correct estimation of nucleosome number. In the Grand Canonical Ensemble, the ensemble-averaged nucleosome number in the entire genome is (supplementary material, section 1.1.3),

Equation (2)

If N were known, the theoretically correct estimate of n1 would be

Equation (3)

such that equation (2) is automatically satisfied. Since $\frac{1}{\max_j{\tilde{O}(\,j)}}$ is not necessarily equal to $\frac{N}{\sum_{j = 1}^L \tilde{n}_1(\,j)}$ , however, the normalization in equation (1) does not guarantee the condition in equation (2). In fact, the nucleosome number estimated by equation (1) is solely determined by the total sequencing depth and the denominator $\max_j{\tilde{O}(\,j)}$ , which may be sensitive to experimental noise and bias.

The incorrect estimation of nucleosome number using equation (1) may confound the inference of nucleosome-positioning energy u and the subsequent analysis of its sequence dependence. We illustrate this issue using a simple simulation, where the nucleosome-positioning energy is fixed to be proportional to the G+C content (GC) of occupied nucleosomal sequence; that is, $u(i) = \beta_{\rm G/C} \sum_{j = 0}^{146} (I_{{\rm C}, i+j} + I_{{\rm G}, i+j}) + u_0 $ , where $\beta_{\rm G/C} = -0.2$ , u0  =  10, and $I_{\alpha, i}$ is the indicator function for nucleotide α at location i. Figure 1 shows the simulation results using the genomic sequence on chromosome I (chrI) of S. cerevisiae. Since energy $u(i)$ is completely determined by the DNA sequence and known, n1 and O can be calculated by solving the direct problem (figure 1(a)). The nucleosome number on chrI is then calculated to be $N = \sum_{i}n_1(i) = 1276$ . An array of well-positioned nucleosomes is evident between the two energy barriers at the locus shown in figure 1(a), reminiscent of statistical positioning [9, 10, 18, 19]. Conversely, assuming that the true n1 and O are known, energy u can be calculated by solving the inverse problem (figure 1(b)). We then fit a linear model of energy that depends only on GC of nucleosome-occupied sequence. As expected, the learned parameters are exactly the same as the true values, $\hat{\beta}_{\rm G/C} = -0.200 $ and $\hat{u}_0 = 10.0$ , with R2  =  1.00; consequently, the predicted $\hat{n}_1$ and $\hat{O}$ obtained from this estimated energy function reproduces the true n1 and O (figure 1(b)). Next, to simulate the incorrect estimation of nucleosome number by equation (1), we have scaled n1 down globally such that the estimated nucleosome number is $\hat{N} = \sum_i n_1(i) = 500$ . The energy calculated from this incorrectly normalized n1 significantly deviates from the true energy (figure 1(c); see figure 1(a)). In this case, a linear model depending only on GC cannot fully capture the incorrectly learned energy (R2  =  0.186), and the fitted parameters, $\hat{\beta}_{{\rm G/C}} = -0.0236 $ and $\hat{u}_0 = 6.79$ , are also off from the true values. Furthermore, the $\hat{n}_1$ and $\hat{O}$ profiles predicted from the fitted linear energy function do not agree with the true n1 and O. In particular, the predicted profiles in figure 1(c) lack a nucleosome array between the two energy barriers present in the true data shown in figure 1(a). Therefore, incorrect normalization of n1 can easily lead to incorrect inference of nucleosome-positioning energy and incorrect predictions of nucleosome profiles; and, conclusions about the sequence-dependence of nucleosome landscape drawn from analyses using this normalization scheme may be confounded.

Figure 1.

Figure 1. Incorrect estimation of nucleosome number may distort the inference of nucleosome-positioning energy. (a) True energy (top panel), n1 (middle panel), and occupancy O (bottom panel) in the simulation. (b) Using the true n1 and O from (a), nucleosome-positioning energy is calculated by solving the inverse problem (orange curve in top panel, hidden by the overlapping green curve). By fitting this calculated energy to a linear sequence model that depends only on GC, the positioning energy can be predicted based on sequence (green curve in top panel). $\hat{n}_1$ (middle panel) and $\hat{O}$ (bottom panel) are predictions from this fitted linear energy model. (c) Same as (b) except that the true n1 is scaled down, so that the nucleosome number is now only 500 (see main text).

Standard image High-resolution image

It is important to note that the above simulation is not an intentional exaggeration of the potential problem. On the contrary, according our investigation, using equation (1) on various in vivo nucleosome maps in S. cerevisiae yields an estimated nucleosome number of at most  ∼30 000 (table S1 of supplementary material), whereas, as shown in the Results section (section 3), the correctly estimated nucleosome number is more than $60\, 000$ , which is close to the common belief that about $80\%$ of the yeast genome is occupied by nucleosomes in vivo [31]. Thus, the normalization in equation (1) underestimates the nucleosome number by half in real in vivo data sets, close to the scenario depicted in figure 1.

Finally, we point out that equation (3), despite being justified as a maximum likelihood estimate, has its own problems. First, the precise nucleosome number N is unknown, and estimating N would require some nucleosome calling algorithm which usually depends on an arbitrary choice of thresholds. Another issue is that the estimated n1 and O are not guaranteed to be between 0 and 1, leading to difficulties in downstream analysis, such as solving the inverse problem to obtain u. To avoid these issues, we will develop in the following section a method that learns N and u simultaneously from experimental data, without requiring a direct estimate of n1 beforehand.

2.2. A cross-entropy method for simultaneously learning nucleosome number and nucleosome-positioning energy

The above difficulty faced by the Locke method (LM) arises because solving the inverse problem of inferring the nucleosome-positioning energy u requires estimating n1 and O as the first step. In order to get around this difficulty, we avoid solving the inverse problem and instead learn u through solving the following optimization problem (supplementary material, section 1.3),

Equation (4)

where $\tilde{O}$ denotes the observed nucleosome occupancy from raw data as previously defined, $O(v)$ the nucleosome occupancy predicted from nucleosome-positioning energy $v$ , and $\mathcal{L}$ a loss function that measures the difference between prediction $O(v)$ and observed $\tilde{O}$ . Throughout the paper, we use a weighted version of negative Pearson correlation coefficients as the loss function (see below). The obtained nucleosome-positioning energy u thus predicts a nucleosome occupancy profile that best correlates with the observed one. Since the Pearson correlation coefficient is invariant under linear transformations of its arguments, no scaling or normalization of $\tilde{O}$ is needed beforehand, and the difficulty described in the previous section thus completely disappears. Importantly, the information of nucleosome number is implicitly contained in the overall shape of $\tilde{O}$ , and our method is able to automatically adjust the nucleosome number to match the observed nucleosome occupancy profile. We will show in the Results section (section 3) that this method can indeed learn the correct nucleosome number empirically observed in high-throughput nucleosome maps. Since the Pearson correlation coefficient can be inflated by extreme outliers, we filter out outlier regions based on observed nucleosome occupancy, resulting in a disjoint set of 'good regions' along the genome (supplementary material, section 1.3)5. Negative average Pearson correlation coefficient across these good regions weighted by the region lengths is then used as the final loss function (supplementary material, section 1.3).

Naively solving the optimization problem in equation (4) is prohibitive, since the number of parameters is of order of the genome length, with $u(i)$ at each genomic location i being a free parameter. In order to reduce the parameter space, we enforce a linear sequence-dependent model:

Equation (5)

where $I_{\alpha, i+j}$ again denotes the indicator function of nucleotide α at genomic location i  +  j, $\beta_{\alpha, j}$ the linear coefficient characterizing how much nucleotide α at position j relative to the start of nucleosome contributes to energy $u(i)$ , and u0 the sequence-independent offset. Note that this paper focuses on mono-nucleotide models, as in equation (5), but a generalization to k-mer models with k  >  1 is straightforward. Using this linear sequence-dependent model, the number of free parameters is only several hundreds when a is set to the canonical nucleosome size 147.

To further reduce the number of parameters, we use a template-based model for the linear coefficient $\beta_{\alpha, j}$ ,

Equation (6)

where $\hat{\beta}_{\alpha, j}$ denotes a pre-defined template of nucleotide α, and $\sigma_{\alpha}$ and $ \mu_{\alpha}$ are free parameters modulating the standard deviation and mean of the template, respectively. The template $\hat{\beta}_{\alpha, j}$ may represent our prior belief about how much different nucleotides at different relative positions contribute to the nucleosome-positioning energy. We construct this template from the negative average nucleotide frequency of nucleosomal sequences obtained from raw sequencing data (supplementary material, section 1.3). Briefly, let $-\tilde{\beta}_{\alpha, j}$ denote the average frequency of nucleotide α at position j relative to the start of the nucleosome, calculated from identified nucleosomes. The larger the value of $-\tilde{\beta}_{\alpha, j}$ , the stronger the nucleosome preference for nucleotide α at relative position j; thus, nucleotide α at j should empirically contribute more negative energy, justifying our choice. For each nucleotide α, we then standardize $\tilde{\beta}_{\alpha, j}$ to obtain the template $\hat{\beta}_{\alpha, j}$ that has mean 0 and standard deviation 1 with respect to j. During optimization, $\sigma_{\alpha}$ in equation (6) are constrained to be nonnegative.

Note that simple nucleosome calling algorithms can be used to obtain $\tilde{\beta}_{\alpha, j}$ from the raw sequencing data, without worrying about estimating the nucleosome number accurately, since we only need the average nucleotide frequency (supplementary material, section 1.3). Furthermore, we enforce reverse-complement symmetry in the model by requiring $\hat{\beta}_{{\rm A}, j} = \hat{\beta}_{{\rm T}, a - 1 - j}$ , $\hat{\beta}_{{\rm C}, j} = \hat{\beta}_{{\rm G}, a - 1 - j}$ , $\sigma_{{\rm A}} = \sigma_{\rm T} := \sigma_{\rm A/T}$ , $\sigma_{\rm C} = \sigma_{\rm G} := \sigma_{\rm C/G}$ , $\mu_{\rm A} = \mu_{\rm T} := \mu_{\rm A/T}$ , and $\mu_{\rm C} = \mu_{\rm G} := \mu_{\rm C/G}$ (supplementary material, section 1.3), such that the energy $u(i)$ calculated from any nucleotide sequence is the same as that calculated from its reverse-complement. Finally, it is easy to see that only one of $\mu_{\rm A/T}$ and $\mu_{\rm C/G}$ is independent, due to the constraint $\sum_{\alpha \in \{{\rm A, C, G, T}\}} I_{\alpha, j} = 1, \forall j$ (supplementary material, section 1.3), allowing us to set $\mu_{\rm A/T} = 0$ . The resulting sequence-dependent energy model is thus

Equation (7)

The motivation behind the choice of model name is as follows. Note that there are in total only four free parameters in the model above: $\sigma_{\rm A/T}$ , $\sigma_{\rm C/G}$ , $\mu_{\rm C/G}$ , and u0. Solving the optimization problem in equation (4) amounts to finding the optimal values of these four parameters, such that the predicted occupancy has the highest correlation with the observed one. $\sigma_{\rm A/T}$ and $\sigma_{\rm C/G}$ characterize how the spatially resolved (SR, as dubbed in Locke et al) pattern of the template $\hat{\beta}_{\alpha, j}$ modulates the positioning energy; similarly, $\mu_{\rm C/G}$ captures the energy dependence on G+C content (GC) of the underlying sequence, thus the name Model GC+SR. u0 effectively modulates the chemical potential of the system and, thus, the nucleosome number (recall that u denotes $\frac{u - \mu}{k_B T}$ throughout the paper).

In order to study how different sequence features, especially GC, 10.5 bp periodicity, and polyA, regulate nucleosome positioning, we constructed two other models with different complexities: Model GC, where the energy depends on GC only, given by

Equation (8)

and Model GC+SR+polyA, where all three components contribute to the energy, given by

Equation (9)

where $m_{\rm polyA}(i)$ denotes the number of bps belonging to a poly(dA:dT) tract (defined as a homopolymer of A or T of length $\geqslant 5$ ) from i to i  −  a  +  1, and $\mu_{\rm polyA} $ characterizes the energy dependence on poly(dA:dT) content. Note that the 10.5 bp periodicity is characterized as part of the spatially resolved sequence motif in $\hat{\beta}_{\alpha, j}$ . The three models Model GC, Model GC+SR, and Model GC+SR+polyA constitute a hierarchy of sequence models, along the same line as in Locke et al, allowing us to dissect the contributions from GC, SR, and polyA to nucleosome positioning, respectively.

To optimize the cost function, we apply the cross-entropy method [3235] (supplementary material, section 1.3). We note that similar ideas of learning the nucleosome energetics through optimization have been used in previous studies [15, 25, 26]. However, [15] mainly focuses on how dinucleotide periodicity affects nucleosome positions in synthetic DNA sequences and genome-wide nucleosome occupancy, and it does not analyze other sequence features such as GC and polyA; it also does not address genome-wide nucleosome positioning; finally, the proposed energy model uses artificial sinusoidal functions. Both [25] and [26] mostly focus on the effect of nucleosome unwrapping. In particular, [25] does not study the sequence-dependence of nucleosome occupancy and positioning. The sequence-dependent nucleosome-positioning energy in [26] is obtained by solving the inverse problem as in Locke et al and thus has the same normalization problem described above. By contrast, our proposed method involves no artificially constructed components and provides a flexible framework that can incorporate several distinct biological features in the energy model, as illustrated by the hierarchical models in equations (7)–(9). Most importantly, our method automatically learns the nucleosome number together with the nucleosome-positioning energy, as will be demonstrated in the following Results sections (section 3).

3. Results

3.1. The cross-entropy method correctly learns the nucleosome number and improves the prediction of nucleosome occupancy

We applied the cross-entropy method (CEM) (section 2; supplementary material, section 1.3) to three independent in vitro nucleosome maps in S. cerevisiae (supplementary material, section 1.4): two using salt dialysis followed by MNase-seq (Kaplan-MNase-invitro-salt [3] and Zhang-MNase-invitro-salt [4]) and another using ACF assembly followed by MNase-seq (Zhang-MNase-invitro-ACF [4]). ACF is an ATP-dependent chromatin assembly factor in Drosophila melanogaster. Model GC+SR (equation (7)) was trained on the three datasets separately, with the templates $\hat{\beta}_{\alpha, j}$ calculated from identified nucleosomes in each dataset. From the learned parameters (table S2 of supplementary material), we then predicted nucleosome occupancy O and nucleosome number $N = \sum_i n_1(i)$ (table 1) by solving the direct problem (section 2; supplementary material, section 1.1). To compare with Locke et al method (LM) [16], we first estimated n1 via equation (1), calculated the nucleosome-positioning energy u by solving the inverse problem (section 2; supplementary material, section 1.1), fitted a linear model as in equation (5) (section 2; supplementary material, section 1.2), and finally predicted nucleosome occupancy O and nucleosome number N (table 1) using the fitted sequence model.

Table 1. Predicted nucleosome number (N) and Pearson correlation coefficient (r) between observed and predicted nucleosome occupancy on three in vitro datasets. As previously described, r was calculated within each 'good region' after filtering out outlier regions (supplementary material, section 1.5). Table shows the average r weighted by region lengths.

    Zhang-MNase-invitro-ACF Kaplan-MNase-invitro-salt Zhang-MNase-invitro-salt
LM r 0.675 0.723 0.566
N 49 372 35 575 52 680
CEM r 0.777 0.732 0.595
N 22 130 29 060 43 280

Since both calculations using CEM and LM accounted for contributions from GC and SR, a difference in their performance is not attributable to different sequence features used by these approaches. Nucleosome occupancy predicted by both methods highly correlated with the observed one (table 1). However, CEM achieved better correlation in all three datasets, despite the fact that the LM linear model (equation (5)) contains many more free parameters than the CEM Model GC+SR (equation (7)). In particular, CEM improved the correlation in Zhang-MNase-invitro-ACF the most, by about 0.1. Interestingly, the predicted nucleosome number of LM differs from that of CEM the most in the same dataset (table 1), indicating that an incorrectly estimated nucleosome number might influence the performance of LM (section 2). Indeed, by simply tuning the nucleosome number (or equivalently, u0 in equation (5)), while keeping the learned energy from LM unchanged, we could achieve better correlations between the predicted and observed nucleosome occupancy in all three datasets (figure S1 of supplementary material). By contrast, the solutions found by CEM already achieved the maximum correlations (figure S1 of supplementary material). Moreover, the maximum correlations obtained by tuning the nucleosome number in LM models were still lower than the correlations achieved by CEM (figure S1 of supplementary material). These results suggest that CEM is indeed learning the correct nucleosome number and demonstrate that it improves the prediction of nucleosome occupancy compared to LM.

We plotted the predicted and observed nucleosome occupancy at the GAL10-1 locus (figure 2)6 and also aligned them at TSS/TTS (figure S2 of supplementary material) for all three in vitro datasets analyzed. Since larger nucleosome number means more tightly packed nucleosomes, we could empirically order the nucleosome numbers in these three datasets from the overall shape of the observed nucleosome occupancy (black curves in figures 2 and S2 of supplementary material). Specifically, Zhang-MNase-invitro-salt showed relatively tightly packed nucleosome arrays within the gene bodies of GAL10 and GAL1 (black curve in figure 2(c)) and around TSS/TTS (black curve in figures S2E and F of supplementary material), while the nucleosome arrays were less pronounced in Kaplan-MNase-invitro-salt (black curves in figures 2(b) and S2C, D of supplementary material) and hardly visible in Zhang-MNase-invitro-ACF (black curves in figures 2(a) and S2A, B of supplementary material). These observations thus suggested that the nucleosome number was the largest in Zhang-MNase-invitro-salt, modest in Kaplan-MNase-invitro-salt, and the smallest in Zhang-MNase-invitro-ACF. Indeed, a quantitative estimation of nucleosome number by comparing the Fourier transform of predicted and observed nucleosome occupancies in the frequency domain confirmed these empirical arguments (figure S3 of supplementary material).

Figure 2.

Figure 2. An example of a genomic locus illustrating that CEM nucleosome occupancy predictions better correlate with the observed profiles compared to LM predictions. (a) Observed (black curve), LM predicted (red curve), and CEM predicted (green curve) nucleosome occupancy for Zhang-MNase-invitro-ACF. (b) Same as (a), but for Kaplan-MNase-invitro-salt. (c) Same as (a), but for Zhang-MNase-invitro-salt. Each curve was standardized by subtracting the mean and then dividing by the standard deviation within the shown region.

Standard image High-resolution image

The nucleosome numbers predicted by CEM were consistent with this ordering, while those predicted by LM were not (table 1), further demonstrating the ability of CEM to better learn the correct nucleosome number. In particular, LM predicted a much larger nucleosome number than CEM for Zhang-MNase-invitro-ACF, and the predicted nucleosome number $49\, 372$ was actually closer to the number $43\, 280$ predicted by CEM for Zhang-MNase-invitro-salt. As a result, the nucleosome occupancy profile predicted by LM in Zhang-MNase-invitro-ACF (red curves in figures 2(a) and S2A, B of supplementary material) showed evident nucleosome arrays and were more similar to the observed nucleosome occupancy in Zhang-MNase-invitro-salt (black curves in figures 2(c) and S2E, F of supplementary material). Interestingly, Kaplan-MNase-invitro-salt used 0.4:1 histone-to-DNA ratio in salt dialysis [3], while Zhang-MNase-invitro-salt used 1:1 [4], indicating that a higher histone-to-DNA ratio indeed results in a larger nucleosome number in vitro. We also note that the data in figures 2 and S2 of supplementary material were standardized to have mean 0 and standard deviation 1, in order to facilitate the visual comparison of correlation; but, the outperformance of CEM was even more evident when normalization (dividing by the genome-wide mean) was used instead (figures S4 and S5 of supplementary material), in which case LM showed only a limited dynamic range in occupancy. Models involving only GC yielded similar comparison results, indicating that our conclusions did not depend on the specific sequence model employed (table S3 of supplementary material, figures S6 and S7 of supplementary material).

Lastly, we note that the predicted nucleosome occupancy in Locke et al manuscript did not show any clear nucleosome array in Zhang-MNase-invitro-salt, neither at the GAL10-1 locus (red curve in supplementary figure S1D of [16]) nor in the average profiles aligned at TSS/TTS (black dashed curves in supplementary figure S7B of [16]). By contrast, our implementations of LM did exhibit evident nucleosome arrays (red curve in figures 2(c) and S2E, F of supplementary material). These discrepancies might have been caused by subtle differences in data processing affecting the normalization factor in equation (1).

3.2. Contributions of G+C content, spatially resolved sequence motifs, and poly(dA:dT) tracts to in vitro nucleosome occupancy

Diverse sequence features, such as GC [36], 10.5 bp periodicity [29, 37], and polyA [38], have been proposed to influence nucleosome landscape. However, which sequence feature actually contributes to which aspect of nucleosome landscape and the degree of these contributions are still under active debate. The flexibility of our CEM framework allows different sequence models to be studied on the same dataset (equations (7)–(9)), providing an opportunity to dissect the complex relationship between different sequence features and different aspects of nucleosome landscape. This section quantifies how much the key sequence features that have received much attention in the community influence nucleosome occupancy.

We trained the aforementioned models in hierarchical order— Models GC, GC+SR, and GC+SR+polyA—on the three in vitro nucleosome maps of S. cerevisiae studied in the previous section (table S4 of supplementary material). To facilitate comparison between different datasets, we used a common template $\hat{\beta}_{\alpha, j}$ for all three datasets. This common template was constructed by combining average nucleotide frequencies of nucleosomes detected by MNase-seq [39] and chemical cleavage [40]; combining these two types of data help eliminate potential MNase digestion biases towards the ends of the nucleosome as well as chemical cleavage biases around the nucleosome dyad (figure S8 of supplementary material). For all datasets considered, the three learned models had negative values of $\mu_{\rm C/G}$ and positive values of $\mu_{\rm polyA}$ (table S4 of supplementary material), supporting the previous hypotheses that GC facilitates and polyA hinders nucleosome occupancy [36, 38]. To compare the relative contributions of the distinct sequence features to nucleosome occupancy, we calculated the Pearson correlation between the genome-wide observed and predicted nucleosome occupancy profiles (table 2). Model GC achieved high correlation in all three datasets, and Model GC+SR showed essentially the same correlations as Model GC. These results suggest that nucleosome occupancy is predominantly determined by GC, while SR, including the 10.5 bp periodicity, does not play a significant role in determining nucleosome occupancy, consistent with previous studies [16, 29]. Furthermore, Model GC+SR+ polyA showed slightly improved correlations in Kaplan-MNase-invitro-salt and Zhang-MNase-invitro-salt, but not in Zhang-MNase-invitro-ACF. To investigate whether polyA really does provide additional information about nucleosome occupancy on top of GC and SR, we calculated the partial correlation coefficient between the observed and Model GC+SR+polyA predicted nucleosome occupancy conditioned on the Model GC+SR prediction (table 2) (supplementary material, section 1.6) [41, 42]. The calculated partial correlation was around 0.2 in both Kaplan-MNase-invitro-salt and Zhang-MNase-invitro-salt, but negligible in Zhang-MNase-invitro-ACF. By contrast, similar partial correlations for Model GC+SR conditioned on Model GC was  <0.1 for all three datasets (table 2). Therefore, for in vitro nucleosome maps obtained from salt dialysis, polyA does make a small contribution to determining genome-wide nucleosome occupancy, while the contribution from SR is negligible.

Table 2. Pearson and partial correlation analyses of CEM nucleosome occupancy predictions for three in vitro datasets. r(data, prediction): Pearson correlation coefficients between observed and predicted nucleosome occupancy from Model GC, GC+SR, and GC+SR+polyA. The first row of partial correlation shows the partial correlation between observed and Model GC+SR-predicted nucleosome occupancy conditioned on Model GC-predicted nucleosome occupancy, representing any additional contribution from SR. The second row of partial correlation shows the partial correlation between observed and Model GC+SR+polyA-predicted nucleosome occupancy conditioned on Model GC+SR-predicted nucleosome occupancy, representing any additional contribution from polyA. All Pearson and partial correlation coefficients were calculated within 'good regions' after filtering out outlier regions (supplementary material, section 1.5) and weighted by the region lengths in the reported average.

    Zhang-MNase-invitro-ACF Kaplan-MNase-invitro-salt Zhang-MNase-invitro-salt
r(data, prediction) Model GC 0.775 0.729 0.589
Model GC+SR 0.775 0.730 0.592
Model GC+SR+polyA 0.775 0.755 0.606
Partial correlation SR 0.048 0.036 0.081
polyA 0.018 0.287 0.161

Locally, it is well known that nucleosome occupancy tends to be reduced around TSS and TTS, resulting in nucleosome depleted regions (NDR) [31]. To study which sequence features predominantly contribute to the formation of NDR, we plotted the nucleosome-positioning energies at TSS and TTS from the three separate components of GC, SR, and polyA in the learned Model GC+SR+polyA (figures 3 and S9 of supplementary material). Again, GC was the primary component responsible for the energy barrier at both TSS and TTS (figures 3(a), (b) and S9A, B, E, F of supplementary material). Compared with GC, polyA showed detectable but smaller energy barriers in datasets obtained from salt dialysis (figures 3(a), (b) and S9A, B of supplementary material), but not in Zhang-MNase-invitro-ACF (figures S9E and F of supplementary material). In comparison, SR showed much suppressed energy barriers. Interestingly, CEM predicted higher energy barriers and lower nucleosome occupancies at TTS than at TSS in all three datasets, consistent with the observed nucleosome occupancy. Finally, unlike in datasets obtained from salt dialysis, neither polyA nor SR showed any influence on nucleosome occupancy in Zhang-MNase-invitro-ACF, suggesting that chromatin assembly factors such as ACF might overwrite certain intrinsic sequence preferences of nucleosome formation.

Figure 3.

Figure 3. Contributions from GC, SR, and polyA in shaping nucleosome occupancy at TSS and TTS. (a) Nucleosome-positioning energy in Model GC+ SR+polyA attributable to GC (yellow curve), SR (green curve), and polyA (red curve) aligned and averaged at TSS of all genes. The total energy is shown in blue. Each energy component is subtracted by its genome-wide mean shown in the legends to facilitate visualization. (b) Same as (a), but aligned at TTS. (c) Observed (blue curve) and predicted nucleosome occupancy from Models GC (yellow curve), GC+SR (green curve), and GC+SR+polyA (red curve), aligned and averaged at TSS and normalized by the genome-wide mean. Pearson correlation coefficients between observation and prediction are shown in the legends. (d) Same as (c), but aligned at TTS.

Standard image High-resolution image

3.3. The dependence of MNase-derived nucleosome occupancy maps on G+C content is substantially biased by MNase digestion

We have shown in the previous section that GC is the primary determinant of the nucleosome occupancy derived from MNase-seq experiments. However, MNase has a substantial GC bias in its cleavage preference [27, 28]. To further investigate the GC dependence of nucleosome occupancy, we calculated the Pearson correlation between GC, polyA, and observed nucleosome occupancy generated by an independent chemical cleavage method that did not involve MNase (Brogaard-chemical-invivo-WT) [40] (figure 4(a)). Since the chemical cleavage data were obtained in vivo, we analyzed an in vivo dataset generated by MNase-seq (Kaplan-MNase-invivo-WT) [3] for comparison (figure 4(d)). As before, chemical-cleavage-derived occupancy showed positive correlation with GC and negative correlation with polyA (figure 4(a)). However, unlike MNase-derived occupancy which correlated, either positively or negatively, most strongly with GC (figure 4(d)), we found that chemical-cleavage-derived occupancy correlated most strongly with polyA (figure 4(a)). Furthermore, partial correlation analysis showed that there was little direct correlation between GC and chemical-cleavage-derived occupancy after conditioning on polyA (figure 4(b)), while the correlation between polyA and chemical-cleavage-derived occupancy was mostly retained even after conditioning on GC (figure 4(b)). This result shows that the marginal correlation between GC and chemical-cleavage-derived occupancy was largely a secondary consequence of the negative correlation of both variables with polyA. By contrast, the GC dependence of MNase-derived nucleosome occupancy was almost unchanged after conditioning on polyA (figure 4(e)). Other independent MNase-derived in vivo and in vitro datasets showed similar trends (figure S10 of supplementary material). Therefore, the GC dependence of nucleosome occupancy was specific to MNase-derived datasets, indicating that this GC dependence might be an artifact of MNase digestion biases.

Figure 4.

Figure 4. The dependence of MNase-derived nucleosome occupancy on G+C content is substantially biased by MNase digestion. (a) Distribution of pairwise Pearson correlation coefficients between chemical-cleavage-derived nucleosome occupancy, GC, and polyA, calculated on 1000 bp intervals tiling the 'good regions'. (b) Distribution of pairwise partial correlation coefficients between chemical-cleavage-derived nucleosome occupancy, GC, and polyA, conditioning on the third variable, calculated on 1000 bp intervals tiling the 'good regions'. (c) A heatmap of nucleosome occupancy as a function of GC and polyA. The S. cerevisiae genome was divided into 1000 bp segments, and each segment was then assigned to a 2-dimensional bin of given GC and poly(dA:dT) content. Color indicates the average nucleosome occupancy in each bin. (d)–(f) Same as ((a)–(c)), but for MNase-derived nucleosome occupancy. (g) Median of the Pearson correlation coefficients of GC and polyA with MNase-derived nucleosome occupancy at different digestion levels, calculated on 1000 bp intervals tiling the 'good regions'. Linear extrapolation was performed to infer the correlation coefficient at digestion time 0. (h) Same as (g), but for partial correlation coefficients.

Standard image High-resolution image

To further resolve the experiment-specific relationship between nucleosome occupancy, GC, and polyA, we divided the genome into 1000 bp segments and binned the segments according to their GC and poly(dA:dT) content. We then plotted the average nucleosome occupancy in the bins as a heat map (figure 4(c) for chemical cleavage occupancy, figure 4(f) for MNase). The existence of a left-to-right gradient in figure 4(c) demonstrated the direct correlation between polyA and chemical-cleavage-derived nucleosome occupancy, while the lack of a top-to-bottom gradient in figure 4(c) showed that there was little direct correlation between GC and chemical-cleavage-derived nucleosome occupancy. This figure also confirmed that it was the negative correlation between GC and polyA that gave rise to the spurious correlation between GC and chemical-cleavage-derived nucleosome occupancy. By contrast, for the MNase-derived dataset, both GC and polyA showed direct correlation with nucleosome occupancy, illustrated by the existence of both weaker left-to-right and stronger top-to-bottom gradient in figure 4(f). Analysis of other independent MNase-derived datasets yielded similar results (figure S10 of supplementary material). These results thus further confirmed that the GC dependence of nucleosome occupancy was specific to MNase-derived datasets.

We reasoned that if the GC dependence of MNase-derived nucleosome occupancy was truly a consequence of MNase digestion biases, then the correlation between GC and nucleosome occupancy should depend on the MNase digestion level. That is, the higher the digestion level, the stronger should be the correlation between GC and nucleosome occupancy. To test this hypothesis, we analyzed a dataset where MNase digestion was performed for 2.5, 5, 10, and 20 minutes [43] (figures 4(g) and (h)). As expected, the correlation between GC and nucleosome occupancy increased with digestion time (figure 4(g)), as did the partial correlation between GC and occupancy conditioned on polyA (figure 4(h)).

Interestingly, the Pearson and partial correlation coefficients showed an almost linear relation with the logarithm of digestion time, although there were only four data points. This linear relationship allowed us to extrapolate the correlation coefficients at digestion time 0, which could be thought of as representing the true biological condition without any MNase bias. Strikingly, the partial correlation between GC and nucleosome occupancy almost dropped to 0 at digestion time 0 in this extrapolation analysis (figure 4(h)), suggesting that almost all of the observed GC dependence of MNase-derived nucleosome occupancy was due to MNase digestion biases. By contrast, the partial correlation between polyA and nucleosome occupancy conditioned on GC changed little with digestion time and was largely retained at the extrapolated time 0 (figure 4(h)), further supporting that the polyA dependence of nucleosome occupancy reflected a true biological signal.

3.4. Contributions of G+C content, spatially resolved sequence motif, and poly(dA:dT) tracts to in vitro nucleosome positioning

We now turn to investigating the sequence determinants of translational and rotational nucleosome positioning. We trained the Models GC, GC+SR, and GC+ SR+polyA on in vitro nucleosome maps generated through MNase-seq experiments (table S4 of supplementary material) and predicted nucleosomes genome-wide by progressively calling genomic locations with the largest predicted n1 as nucleosome start positions, while constraining that the called nucleosomes cannot overlap with each other (supplementary material, section 1.5). The nucleosome calls were thus ranked by the value of n1. Similarly, positioned nucleosomes were also identified from experimental data by using the same algorithm. To assess the agreement between observed and predicted nucleosome positions, we examined top N nucleosomes from the respective ranked lists, where N was set to be the nucleosome number predicted by Model GC in each dataset (table S4 of supplementary material; supplementary material, section 1.5). The three models showed modest and similar prediction power on single nucleosome positions (figure 5(a)), being able to predict about 50% of observed nucleosomes within 50 bps. In particular, adding the SR component to nucleosome-positioning energy did not significantly improve the prediction accuracy, agreeing with our previous finding based on spectral analysis that sequence periodicity does not play a major role in translational positioning [29].

Figure 5.

Figure 5. Spatially-resolved sequence motifs, including the 10.5 bp periodicity, facilitate the rotational but not translational positioning of nucleosomes. (a) Cumulative distribution of absolute dyad-to-dyad distance between predicted and observed nucleosome positions. Grey curve represents a random control using non-overlapping uniformly distributed nucleosomes. (b) Distribution of dyad-to-dyad distance between predicted and observed nucleosome positions. (c) Distribution of absolute dyad-to-dyad distance between predicted and observed nucleosome positions mod 10 bp. (d) Distribution of dyad-to-dyad distance between redundant nucleosomes.

Standard image High-resolution image

By contrast, the distribution of dyad-to-dyad distance between the nucleosome positions observed in sequencing data and those predicted by the models including an SR component was enriched for multiples of 10 bp (figure 5(b)), and the rotational nucleosome positions were also better predicted by these models (figure 5(c)). In order to better capture cell-to-cell variation in nucleosome positions, we progressively identified a redundant set of nucleosomes containing five times the above number of nucleosomes in each dataset. The dyad-to-dyad distance between nucleosomes within this redundant set exhibited a 10 bp oscillatory pattern emerging from rotational positioning of nucleosomes (black curve in figure 5(d)). Importantly, this 10 bp oscillation could be recapitulated by a model only if it included the SR component in its nucleosome-positioning energy (figure 5(d)). These results thus demonstrated that spatially-resolved sequence motifs, including the 10.5 bp periodicity, facilitated the rotational positioning of nucleosomes, again confirming our previous conclusion based on spectral analysis [29]. Analysis of other in vitro nucleosome maps provided similar results, except that Zhang-MNase-invitro-ACF showed little sequence dependence other than GC, as previously mentioned (figure S11 of supplementary material).

3.5. Models on in vivo nucleosome maps exhibit a high degree of experiment-specific variability

We have so far focused on in vitro nucleosome maps, since in vitro conditions provide a simplified platform for studying intrinsic sequence preferences hidden in the nucleosome landscape. In vivo, DNA sequence is only one of many factors that can influence nucleosome positioning. Trans-factors, such as chromatin remodelers, transcription factors and RNA polymerase, can actively regulate nucleosome positions. In order to quantify the role of DNA sequence in determining the in vivo nucleosome landscape, we applied CEM to available in vivo nucleosome maps. See supplementary material, section 1.4 for detailed information about these datasets and our labeling convention. CEM-predicted nucleosome occupancy achieved consistently higher correlation with observed nucleosome occupancy than the LM prediction (table S5 of supplementary material), demonstrating that the benefits of CEM generalized to in vivo datasets. However, CEM trained on different in vivo nucleosome maps yielded discordant results (table S5 of supplementary material). In particular, the estimated nucleosome number, although consistently larger than that predicted by LM, varied between  ∼36 000 and  ∼56 000.

Since all curated in vivo nucleosome maps were generated using wild-type S. cerevisiae in log phase, one would expect similar nucleosome numbers in these datasets. Indeed, a quantitative estimation of nucleosome number by matching the Fourier transform of observed nucleosome occupancy profiles showed that all analyzed in vivo nucleosome maps contained a similar number of  ∼65 000 nucleosomes (figure S12 of supplementary material), suggesting that CEM was consistently underestimating the nucleosome number in these datasets (table S5 of supplementary material). Visually comparing predicted and observed nucleosome occupancy profiles showed that, for datasets where nucleosome number was significantly underestimated (Kaplan-MNase-invivo-WT [3] and Ocampo-MNase-invivo-WT [39]), CEM-predicted nucleosome occupancy was indeed missing a large fraction of nucleosomes evident from experimental data (figure S13 of supplementary material). Of note, the observed nucleosome occupancy in these datasets showed peculiar long-range fluctuations which were not observed in McKnight2016-MNase-invivo-WT-log-80 [44] and McKnight2015-MNase-invivo-WT-log-50 [45], where nucleosome number was only modestly underestimated (figure S13 of supplementary material). Interestingly, the latter two datasets, McKnight2016-MNase-invivo-WT-log-80 and McKnight2015-MNase-invivo-WT-log-50, that we were able to model better, were from the same lab and generated by the same author(s). Fourier analysis of the observed nucleosome occupancy further demonstrated that these two datasets also had relatively large spectral density at the peak around  ∼200 bp corresponding to a nucleosome-size signal. By contrast, other in vivo nucleosome maps contained smaller spectral density at the peak around  ∼200 bp, but had more spectral energy in long-range components (figure S12A of supplementary material). These results suggested that experiment-specific noises and biases, such as a trending effect, might be confounding the CEM optimization.

Since the two datasets McKnight2016-MNase-invivo-WT-log-80 and McKnight2015-MNase-invivo-WT-log-50 exhibited the smallest long-range trending effect and resulted in only a modest underestimation of nucleosome number, we used these two datasets to study the in vivo nucleosome landscape by training the aforementioned Models GC, GC+SR, and GC+SR+polyA (tables S6 and S7 of supplementary material). As in the in vitro case, GC was still the primary factor determining in vivo nucleosome occupancy (table S6 of supplementary material), although this GC dependence might have arisen from MNase digestion biases, as previously described. Also similar to the in vitro findings, polyA made a small contribution to determining in vivo nucleosome occupancy (table S6 of supplementary material). One difference was that SR seemed to have a greater impact on nucleosome occupancy in vivo than in vitro (table S6 of supplementary material), especially at the TSS (figure S14 of supplementary material). For single nucleosome positions, the prediction accuracy based on sequence information alone was very modest, although SR still showed signatures of facilitating rotational positioning (figure S15 of supplementary material).

3.6. In vivo nucleosome occupancy around TSS is partially determined by DNA sequence

It has been known for a long time that nucleosome occupancy in vivo tends to be reduced around TSS, leading to a nucleosome depleted region (NDR), and that the average nucleosome occupancy downstream of TSS exhibits an oscillatory pattern reflecting statistical positioning [18, 31]. It has been argued that DNA sequence itself is not sufficient to explain the NDR and oscillatory pattern [16] and that active remodeling by trans-factors are needed [23]. Ostensibly consistent with these claims, the CEM-predicted occupancy using Model GC+SR+polyA, averaged over all genes at TSS, did not show an NDR that was as deep as in the observed occupancy (figures 6(a) and S16A of supplementary material). Furthermore, the oscillations in CEM-predicted average occupancy were very weak (figures 6(a) and S16A of supplementary material), seemingly supporting the idea that DNA sequence alone is not able to fully explain the energy barrier present at TSS.

Figure 6.

Figure 6. In vivo nucleosome occupancy around TSS is partially determined by DNA sequence. Shown are the results using the Model GC+SR+polyA trained on McKnight2016-MNase-invivo-WT-log-80 [44]. (a) Observed (blue) and predicted (yellow) nucleosome occupancy aligned at TSS and averaged over all genes. (b) Genes were ranked by the Pearson correlation coefficient between observed and predicted nucleosome occupancy within $\pm $ 1 kb of TSS and divided into quintiles (different colors). The distribution of these correlation coefficients are shown. (c) Observed (solid curves) and predicted (dashed curves) nucleosome occupancy aligned at TSS and averaged over the genes within each quintile from (b). Average DNase I hypersensitivity [46] (gray curve and shade) is also shown for genes within each quintile. (d) Enrichment analysis for expression variability [47]. All genes are ranked between 0 and 1 according to a variability measure. The median rank of genes within each quintile is shown. A permutation test is performed to assess whether the genes within each quintile are significantly enriched for high or low variability.

Standard image High-resolution image

To investigate the nucleosome occupancy around TSS in greater detail, however, we ranked the genes according to their predictability, as measured by the Pearson correlation between observed and predicted nucleosome occupancy within $\pm $ 1 kb from TSS, and divided the genes into quintiles (figures 6(b) and S16B of supplementary material). For genes in the fifth quintile showing the highest predictability, the NDR depth and oscillatory pattern in predicted occupancy better matched the observed patterns (figures 6(c) and S16C of supplementary material); for this subset of genes, DNA sequence thus played a more significant role in regulating nucleosome occupancy at TSS. Interestingly, this same subset of genes showed less DNase hypersensitivity [46] around TSS (figures 6(c) and S16C of supplementary material), smaller expression variability as measured in [47] (figures 6(d) and S16D of supplementary material), and a slightly lower histone turnover rate [48] (figure S17 of supplementary material). Importantly, the oscillation of average nucleosome occupancy for this class of genes was predicted from DNA sequence alone, using the simplest model accounting for only hard-core interactions between adjacent nucleosomes. In sharp contrast, previous studies either failed to reproduce this oscillatory pattern using sequence alone [16], potentially due to their underestimation of nucleosome number, or required artificially constructed energy barriers at TSS [23]. The fact that DNA sequence alone could partially recapitulate the average nucleosome occupancy in a subset of genes highlighted that different genes might utilize distinct mechanisms to regulate nucleosome formation at TSS.

3.7. In vivo nucleosome occupancy around TTS is primarily determined by DNA sequence

Finally, we investigated the pattern of in vivo nucleosome occupancy around TTS. Unlike at TSS, the Model GC+SR+polyA was able to fully capture the nucleosome depletion at TTS (figures 7(a) and S18A of supplementary material). The slight discrepancy between prediction and observation at TTS could be explained by the existence of a nearby TSS [49]. When sorted by the distance to the closest TSS (figures 7(b) and S18B of supplementary material), the TTS isolated from any TSS had an even better agreement between predicted and observed nucleosome occupancy (figures 7(c) and S18C of supplementary material); by contrast, those TTS close to TSS showed greater deviation from prediction (figures 7(c) and S18C of supplementary material). We note, however, that these observations might be specific to MNase-derived nucleosome maps. For instance, the sharper energy barrier at TTS predicted by CEM was mostly attributable to GC (figure S14 of supplementary material), but according to our previous analysis, the GC dependence of MNase-derived nucleosome maps were subject to substantial MNase digestion biases. Moreover, nucleosome maps obtained from experimental methods avoiding MNase digestion showed much less pronounced depletion at TTS than the maps obtained from MNase-seq [50].

Figure 7.

Figure 7. In vivo nucleosome occupancy around TTS is primarily determined by DNA sequence. Shown are the results using the Model GC+SR+polyA trained on McKnight2016-MNase-invivo-WT-log-80 [44]. (a) Observed (blue) and predicted (yellow) nucleosome occupancy aligned at TTS and averaged over all genes. (b) Genes are ranked by the distance from their TTS to the closest TSS and divided into quintiles (different colors). The distribution of these distances are shown. (c) Observed (solid curves) and predicted (dashed curves) nucleosome occupancy aligned at TTS and averaged over the genes within each quintile from (b).

Standard image High-resolution image

4. Discussion

In this paper, we have developed a cross-entropy method (CEM) for simultaneously learning the nucleosome number and nucleosome-positioning energy from empirical genome-wide nucleosome maps. CEM circumvents a crucial computational difficulty encountered by previous approaches when estimating the nucleosome number [16]. Applying our framework to various in vitro and in vivo datasets in S. cerevisiae has provided a comprehensive quantification of the role of DNA sequence in shaping the genome-wide nucleosome landscape that includes the distinct concepts of nucleosome occupancy and nucleosome positioning.

Using in vitro nucleosome maps, we have shown that spatially resolved sequence patterns, including the controversial 10.5 bp nucleotide periodicity, facilitate rotational, but not translational positioning, consistent with the previous conclusion based on categorical spectral analysis of nucleosomal DNA [29]. In our models, GC is the primary determinant of MNase-derived nucleosome occupancy, while the ostensibly minor role of polyA may be under-estimated due to a confounding effect. More precisely, our partial correlation analysis has demonstrated that the GC dependence of MNase-derived nucleosome occupancy may be largely attributable to MNase digestion biases, in line with a recent deep-learning study showing that GC contributes to the classification of only a small fraction of nucleosomal sequences in S. cerevisiae [51]. However, we cannot rule out the possibility that the intrinsic sequence preferences of MNase digestion and nucleosome formation are partially overlapping. Using currently available sequencing data, it is difficult to dissect how much of the GC dependence is due to MNase bias and how much, if any, is truly nucleosome preference. Further in vitro studies using MNase-independent experiments, such as the chemical cleavage method [40, 52, 53], should improve our understanding in the future. By contrast, the effect of polyA on nucleosome occupancy is robust and independent of the specific experimental method used, suggesting that polyA, together with the steric exclusion between adjacent nucleosomes, may be the truly dominant factor in determining genome-wide nucleosome occupancy.

Applying CEM to in vivo nucleosome maps has shown, for the first time, that the canonical oscillatory pattern in average nucleosome occupancy around TSS can be partially recapitulated by DNA sequence and steric exclusion of nucleosomes for a subset of genes. We believe that previous studies have not been able to capture this pattern, because of the aforementioned underestimation of true nucleosome number [16]. This result also highlights a key difference between in vitro and in vivo nucleosome maps: in vivo conditions tend to yield a larger nucleosome number, or equivalently, higher chemical potential. Although some in vitro experiments have used histone-to-DNA ratios close to those under in vivo conditions [4], histone chaperones and chromatin remodelers may actively package histones onto DNA, effectively increasing the chemical potential in vivo. This difference between in vitro and in vivo conditions, although previously noted [15, 22], has not been fully appreciated to date, especially regarding the formation of regularly spaced nucleosome arrays around TSS.

Specifically, the oscillatory pattern in average nucleosome occupancy around TSS has been thought to be a unique feature of the in vivo nucleosome landscape created by regulatory factors other than DNA sequence itself. However, the fact that DNA sequence alone in our in vivo models can reproduce this oscillatory pattern for select genes suggests that, given high enough chemical potential, this pattern may be also produced in vitro. This hypothesis may be tested by mapping nucleosome positions in vitro with an even higher histone-to-DNA ratio than that used in [4]. In fact, our analysis shows that nucleosome arrays are already vaguely visible in some in vitro nucleosome maps [4] (figures 2(c) and S5E, F of supplementary material) and that tuning up the chemical potential to match the expected in vivo nucleosome number indeed creates a more pronounced oscillatory pattern around TSS (figure S19 of supplementary material). We also note that the effective chemical potential may be different at different genomic locations, especially in vivo [15]. A future study where separate models are trained on distinct genomic segments may help elucidate our hypothesis [15].

Our method represents a unified framework that accounts for both intrinsic sequence preference of nucleosome formation and steric exclusion between neighboring nucleosomes. In order to fully complete the story, other factors such as chromatin remodelers, transcription factors, and RNA polymerase need to be taken into account. One advantage of our computational framework based on CEM is that it can be potentially extended to include these regulatory factors. To a first order approximation, the presence of these trans-factors can be considered as modulating the effective nucleosome-positioning energy in addition to DNA sequence. For example, an additional term that linearly depends on the ChIP-seq intensity of relevant trans-factors can be introduced into equation (9). Another related strategy is to use CEM to study specific functions of individual chromatin remodelers, utilizing the data generated through in vitro genome-wide reconstitution of nucleosomes in the presence of individual chromatin remodelers [54]. For example, it has been shown that RSC clears the promoter by reading poly(dA:dT) as a directional nucleosome removal signal [54]. We expect this effect to be captured by a relative change in $\mu_{\rm polyA}$ in equation (9). Similarly, INO80 has been shown to help position the  +1 nucleosome [54], suggesting that INO80 may act as an additional energy barrier at TSS. Thus, incorporation of INO80 ChIP-seq information into the sequence model may help improve the prediction of nucleosome landscape. Lastly, it is known that ISW1 can tighten the spacing in nucleosome arrays downstream of TSS [54]. A plausible hypothesis is that ISW1 increases the effective chemical potential so that the nucleosomes in the region are more tightly packed. All of these aspects can potentially be captured by CEM with different energy models.

Finally, it is important to note that other statistical mechanics-inspired models that relax the assumption of hard-core interactions have been also used to study nucleosomes. For example, nucleosome unwrapping and breathing can be captured by a soft-core interaction [25] or variable nucleosome size [20, 26]. More complex nearest neighbor interactions imposed by higher-order chromatin structure can be modeled [23, 24]. Models beyond equilibrium statistical mechanics have also been attempted [5557]. Nevertheless, this paper demonstrates that even the simplest model involving only steric exclusion can still provide new insights into the principles that guide nucleosome occupancy and nucleosome positioning.

Acknowledgments

This work was supported by NSF [1442504] and NIH [R01CA163336].

Footnotes

  • See online supplementary material for detailed information of datasets analyzed and additional description of computational methods.

  • We note that a similar procedure is performed in Locke et al [16] to filter out outlier regions.

  • GAL10-1 locus was selected for visualization in order to compare with the results in supplementary figure S1 of Locke et al [16].

Please wait… references are loading.
10.1088/1478-3975/aadad2